Materials Analysis Applications Compendium BR53012
Materials Analysis Applications Compendium BR53012
Materials Analysis Applications Compendium BR53012
NIR A case study of using FT-NIR for pharmaceutical hot melt extrusion process monitoring 14
NMR Measuring the equilibrium constant of a keto-enol tautomerism using benchtop NMR 20
Benchtop NMR combined with GC/MS confirms identity of forensic case sample 23
FT-Raman: an invaluable addition to the forensic arsenal to combat the opioid epidemic 54
A B
Figure 5: Far-IR ATR spectra: (A) images of red pigment extracted from
Figure 4. Qing dynasty lacquer clothes wardrobe and inset shows a wardrobe sample, (B) far-IR ATR spectrum of wardrobe sample and (C)
cross-section from the interior used for analysis. (A) Exterior, (B) interior, vermillion reference.
and (C) cross-section analyzed from the compound wardrobe. The
asterisk represents the original decorative surface. The layers above
this point are later additions. Accession number: 1940-7-2, Philadelphia
Museum of Art; Artist/maker unknown Qing Dynasty (1644–1911);
Lacquered wood with painted and gilt decoration; brass fittings.
Purchased with the Bloomfield Moore Fund, 1940.
2. Meiluna, Raymond J., James G. Bentson, and Arthur Steinberg. Analysis of Aged Paint Binders by FTIR
Spectroscopy. Studies in Conservation 35 (1990).
3. Sloggett, Robyn, Caroline Kyi, Nicole Tse, Mark J. Tobin, Ljiljana Puskar, and Stephen P. Best. Microanalysis of
Artworks: IR Microspectroscopy of Paint Cross-sections. Vibrational Spectroscopy 53.1 (2010): 77-82.
4. Newman, Richard. Some Applications of Infrared Spectroscopy in the Examination of Paintings Materials. Journal
of the American Institute for Conservation 19.1 (1979): 42-62.
5. Manzano, E., N. Navas, R. Checamoreno, L. Rodriguez-Simon, and L.F. Capitan-Vallvey. Preliminary Study of UV
Ageing Process of Proteinaceous Paint Binder by FTIR and Principal Component Analysis. Talanta 77.5 (2009):
1724-731.
6. Jonsson, Julia, and Tom Learner. Separation of Acrylic Paint Components and Their Identification with FTIR
Spectroscopy. Proceedings of the Sixth Infrared and Raman Users Group Conference (IRUG6). Florence, Italy
March 29th-April 1st 2004. Ed. Marcello Picollo. Florence: Il Prato, 2005. 58-68.
7. Hodson, J., and J.A. Lander. The Analysis of Cured Paint Media and a Study of the Weathering of Alkyd Paints by
Fourier Transform Infra-red/photoacoustic Spectroscopy. Polymer 28.2 (1987): 251-56.
8. Lang, Patricia L., Chad D. Keefer, Jessica C. Juenemann, Khoa V. Tran, Scott M. Peters, Nancy M. Huth, and
Alain G. Joyaux. The Infrared Microspectroscopic and Energy Dispersive X-ray Analysis of Paints Removed from a
Painted, Medieval Sculpture of Saint Wolfgang. Microchemical Journal 74.1 (2003): 33-46.
9. Olin, J.S. The Use of Infrared Spectrophotometry in the Examination of Paintings and Ancient Artifacts. Instrument
News 17 (1966): 1.
10. McClure, A., J. Thomson, and J. Tannahill. Infrared Spectra of Ninety-six Organic Pigments. Journal of Oil and
Colour Chemists’ Association 51 (1968)
11. López-Gil, Ruiz-Moreno, Miralles. Optimum acquisition of Raman spectra in pigment analysis with IR laser diode
and plulsed UV irradiation Journal of Raman Spectroscopy. 37 (2006); 966-973
12. Kendix, Elsebeth L., Silvia Prati, Edith Joseph, Giorgia Sciutto, and Rocco Mazzeo. ATR and Transmission
Analysis of Pigments by Means of Far Infrared Spectroscopy. Analytical and Bioanalytical Chemistry 394 (2009):
1023-032.
13. http://lisa.chem.ut.ee/IR_spectra/
14. Afremow, Leonard C., and John T. Vanderberg. High Resolution Spectra of Inorganic Pigments and Extenders in
the Mid-Infrared Region from 1500 cm-1 to 200 cm-1. Journal of Paint Technology 38.495 (1966): 169-202
15. C. Karr and J. Kovach. Far-Infrared Spectroscopy of Minerals and Inorganics. Appl. Spectrosc. 23, 219 (1969).
16. M.R. Nelson, Authentic or Not, ChemMatters, April 2011, pp. 15-17.
In this note, we demonstrate the use of an imaging ATR accessory that’s well
suited for the analysis of these samples. The accessory makes contact with the
sample just once, while still permitting measurements across the sample surface.
Additionally, germanium ATR microscope measurements offer the benefit of a
four times magnification due to presence of germanium (refractive index n=4)
instead of air (n=1) at the sample interface, resulting in enhanced spatial resolution
compared to non-ATR measurements.
Experimental
Spectra were measured using a Nicolet iN10 MX infrared imaging microscope
(Figure 1), configured with the following detectors: a room temperature DTGS, a
single point MCT-A and a linear array MCT-A detector. Each sample cross section
was cut to a size of approximately 5 x 5 x 2 mm (l x w x h) and was placed on the
sample post of an imaging ATR microscope accessory (Figure 1) and raised to
make contact with the germanium ATR crystal for measurement.
Unless otherwise indicated, spectra were collected using
the array detector in “ultra-fast” mode where a single scan
per step at 16 cm-1 resolution was collected at each map
point at a spatial resolution of 6.25 microns. Sample and
background maps of about 625 x 625 microns consisting of
about 11,000 spectra were measured in approximately 1.5
minutes each.
Results
Antacid product component distribution
Representative map spectra and their comparison to the Figure 2: Maps showing distribution of main components of
library spectra of pure materials are shown in Figure 2. antacid product (top) and representative spectra shown in red,
Examination of individual map spectra showed that compared to library spectra shown in blue (bottom)
Author Abstract
Suja Sukumaran Fourier-transform infrared (FTIR) spectroscopy is one of the most
Thermo Fisher Scientific, USA versatile analytical tools used across various disciplines. In this study, the
Thermo Scientific™ Nicolet™ iS10 and Nicolet iS50 FTIR spectrometers,
Keywords equipped with Attenuated Total Reflection (ATR)-FTIR and Transmission-
FTIR, ATR, protein structure FTIR, were used for the determination of protein secondary structures.
elucidation, Biocell calcium Structure calculations based on a protein database as well as spectral
fluoride cell, ConcentrateIR2 ATR, deconvolution are discussed. The analyses were quick and easy.
transmission
Introduction
Protein secondary structure describes the repetitive conformations of
proteins and peptides. There are two major forms of secondary structure,
the α-helix and β-sheet, so named by the patterns of hydrogen bonds
between amine hydrogen and carbonyl oxygen atoms that create the
peptide backbone of a protein.1 Understanding protein secondary structure is
important to gain insight into protein conformation and stability. For example,
temperature dependent analysis of the secondary structure is critical in
determining storage conditions for maintaining active therapeutic proteins.2
Protein secondary structure is also crucial in understanding the structure-
function relationship and enzyme kinetics of various proteins.3
In this application note, we demonstrate the use of FT-NIR for in-line API
concentration monitoring during an HME process. The considerations in deriving
an appropriate partial least square (PLS) model are discussed. In addition, the
feasibility of employing NIR spectral responses as an indicator for HME process
stability is also explored.
Experimental
An 11-mm twin-screw extruder, Thermo Scientific™ Pre-blended ketoprofen (API) and Eudragit® L100-55
Process 11 (Thermo Fisher Scientific, Karlsruhe, (polymer), with their chemical structures shown in Figure
Germany) was used for the HME processes. Figure 1 2, were fed into the extruder at Zone 1 position. Zones 2
is a schematic showing the screw profile design as well through 8 were mixing, conveying and discharge zones.
as process configuration, and the effective length-to- Zone 9 is a die that reduces the barrel configuration to a
diameter ratio (L/D) for the system was defined as 40:1. single circular orifice.
Three mixing zones were configured for the process.
polymer/API
pre-blend
Figure 1: Schematic of a Thermo Scientific Process 11 Twin Screw Extruder configured in this study.
A Thermo Scientific™ Antaris™ II MDS FT-NIR The target blending ratio of ketoprofen and Eudragit
spectrometer (Thermo Fisher Scientific, Madison, L100-55 was 50:50 (w/w). The temperature in the mixing
Wisconsin), equipped with a HME diffuse reflectance and melting zones was set at 120 °C. The material feed
probe, was used for spectral acquisition. The reflectance rate was 100 g/h. NIR data acquisition was performed
probe has standard 1/2-20 UNF mounting threads and using the Thermo Scientific™ RESULT™ v3 software
was screwed into one of the sensor ports of the exit package. The spectral range was 4000–10,000 cm-1
block, between the exhaust gas vent and the exit orifice. and the resolution was 8 cm-1. The number of spectral
Since the hot melt is a translucent liquid at the sampling scans averaged was 16, or 8 seconds per spectrum. A
location, a metal reflector was placed on the opposite background reference spectrum was acquired using the
side of the probe to enhance the NIR signal. The gap transmission sampling module at the beginning of each
between the reflector and the probe surface was ~2 mm. set of experiments.
The exit block was heated at 120 °C by an electric heater
controlled by the extruder PLC.
Results and discussion To build a calibration model to predict the API
Building a PLS model to monitor API concentration during HME, the pre-blend API/polymer
concentration for HME ratio was varied from 40 to 60%, bracketing the
Figure 3A shows the overlaid NIR spectra of ketoprofen, target API concentration of 50% (w/w). The normal
Eudragit, and a 50:50 (w/w) extrudate. Both ketoprofen process temperature was set at 120 °C, but +/- 10 °C
and Eudragit are powders at room temperature, and their disturbances were introduced to simulate possible
absorption spectra were measured using an integration temperature variation in manufacturing process. The
sphere module of the Antaris II spectrometer. The feed rate of 100 g/h remained constant throughout
spectrum of the extrudate was acquired with an NIR the experiment. After each API concentration change
HME probe during an extrusion process. All spectra have was introduced, adequate time was allowed to ensure
strong features in the first overtone C-H stretching region a steady state was reached before the next process
(~6000 cm-1). In addition, both Eudragit and the extrudate change. The spectral data collected during steady states,
have noticeable peak features in the first overtone O-H including those at different temperatures, were used
stretching region (~6900 cm-1). as calibration samples. NIR spectra were collected at
a 20-second interval. The calibration sample set has a
total of 85 spectra at 7 concentration levels: 40%, 45%,
47.5%, 50%, 52.5%, 55%, and 60%.
A
Figure 3B shows the overlaid NIR spectra of the
calibration samples. There is a noticeable baseline up-
drift in all spectra. While it is difficult to ascertain the
exact causes for this up-drift, one likely contributing
factor is the scattering by the air bubbles formed
during the HME process and the variation in effective
pathlength. Because of the relatively low material feed
rate (100 g/h); hence, high air/material ratio, the vent
nozzle was not able to completely remove air and
moisture from the hot melt, resulting in micro-size air and
water vapor bubbles.
B
The Norris second derivative10 was first applied to the
raw spectra to remove the baseline drift, followed by a
standard normal variant (SNV) to minimize the spectral
pathlength variation. A PLS model11 to estimate the
ketoprofen concentration was then derived using the
Thermo Scientific™ TQ Analyst™ software. All spectral
data were mean-centered in the PLS regression, and
the spectral range of 5500 to 6650 cm-1 was used
for the correlation. A total of 85 spectra were divided
Figure 3. (A) Absorption NIR spectra of ketoprofen, Eudragit L100-
55, and a 50/50 (w/w) ketoprofen/Eudragit extrudate in full scale. into two groups: calibration (73 spectra) and validation
(B) Transflectance NIR spectra in common scale acquired during (12 spectra). PLS finds its significant model factors
extrusion processes.
from the Leave-One-Out cross validation. The number
NIR spectra are often complex and contain broad of significant PLS factors represents the number of
overlapping absorption bands. Chemical, physical, independent variables that affect sample spectral
and structural properties of all species present in a responses, such as concentration, impurities, density,
sample influence the measured spectra. As a result, the opaqueness, or sample color. In the current case,
measured NIR spectra depend on more than one variable the developed PLS model (Figure 4) has a correlation
simultaneously; hence, multivariate. Chemometrics uses coefficient of 0.998 and a root mean squared error of
mathematical and statistical procedures, such as PLS calibration (RMSEC) of 0.43%. The root mean squared
and principle component analysis (PCA), for multivariate error of prediction (RMSEP) from the validation sample
data analysis to filter information that correlates to a group is 0.62%.
certain property from a large set of data.
The calibration model was then applied to the spectra
collected during the HME process, and the results
are shown in Figure 5. As can be seen, the predicted
API% tracks the pre-blend API profile. With a feed
rate of 100 g/h, the process has a residence time of
~10 min, but it takes 2-3 multiples of the residence time
(approximately 20-30 min) to establish a new steady state
after the pre-blend ratio is introduced. Figure 4. Calibration results of API measurement.
2. Thiry J., Krier F., Evrard B. A review of pharmaceutical extrusion: Critical process parameters and scaling-up, Int.
J. Pharmaceut., 2015, 479:1, 227-240.
3. Breitenbach J. Melt extrusion: from process to drug delivery technology, Eur. J. Pharm. Biopharm., 2002, 54:2,
107-117.
4. Crowley M.M., Zhang F., Repka M.A., Thumma S., Upadhye S.B., Battu S.K., McGinity J.W., and Martin C.
Pharmaceutical applications of hot-melt extrusion: Part I, Drug Dev. Ind. Pharm., 2007, 33:9, 909-926.
5. FDA Guidance for Industry. PAT – a framework for innovative pharmaceutical development, manufacturing, and
quality assurance. Rockville, MD: Food and Drug Administration, 2004.
6. ICH Q8(R2), Pharmaceutical Development, Part I: Pharmaceutical development, and Part II: Annex to
pharmaceutical development, 2009.
7. Markl D., Wahl P.R., Menezes J.C., Koller D.M., Kavsek B., Francois K., Roblegg E., and Khinast J.G. Supervisory
control system for monitoring a pharmaceutical hot melt extrusion process, AAPS PharmSciTech, 2013, 14:3,
1034-44.
8. MacPhail N., Meyer R.F, Phillips, J.X., Gendron C.M., Smith-Goettler B. NIR Monitoring of a hot-melt extrusion
process, Spectroscopy Special Issues, 2011, 26:8.
9. Abdi H. and Williams L.J., Principal component analysis, Wires. Comput. Stat., 2010, 2, 433-459.
10. Rinnan A., van den Berg F., Engelsen S.B. Review of the most common pre-processing techniques for near-
infrared spectra, Trends Anal. Chem., 2009, 28:10, 1201-1222.
11. Geladi P. and Kowalski B.R., Partial least-squares regression: a tutorial, Anal. Chim. Acta, 1986, 185, 1-17.
No. AN52889
Benchtop NMR Combined with
GC/MS Confirms Identity of Forensic
Case Sample
Authors: Dean Antic, Ph.D.,
Thermo Fisher Scientific,
San Jose, CA, USA
WanLi Wei, Senior Engineer,
Institute of Forensic Sciences, Wuxi Public
Security Bureau, Wuxi, China
Key words
NMR, GC/MS, picoSpin, illicit drug, forensics The case sample demonstrated in this note exemplifies
the need for multiple techniques in order to confirm the
The benchtop Thermo Scientific™ picoSpin™ 80 1H identity of seized forensic samples, as recommend by
NMR spectrometer provides an additional layer of SWGDRUG.
structural identification of drug analogues and precursors,
complementing GC/MS analysis. The combination of Introduction
the two techniques allows for a positive identification of Illicit drugs identification is a challenge to law enforcement
real forensic case samples with high confidence, thereby due to the vast assortment of illegal drugs that already
enhancing the presumptive testing capabilities of illicit exist, and an increasing number of new, “not-yet” illegal
drug screening facilities. analogues of classified drugs appearing on the street.
Added to this burden is the identification of clandestine
Abstract lab chemicals and precursor compounds used in the
A forensic case sample of an illicit drug precursor was manufacture of illicit drugs, as well as the excipients used
analyzed by GC/MS and benchtop NMR. The most as adulterants to alter street-level drug purity.
probable chemical structure from a GC/MS library
search, however, conflicts with the spectral features Law enforcement has a variety of analytical tools at their
identified by NMR. By interrogating the peak pattern and disposal, including color tests, Fourier transform infrared
chemical shifts in the NMR spectra, both experimentally (FTIR), gas chromatography/mass spectrometry (GC/MS),
obtained and predicted by Mnova NMR software, the Raman spectroscopy and nuclear magnetic resonance
seized sample was identified as a compound with a spectroscopy (NMR) to aid in characterizing forensic
lower matching score by GC/MS. The combination of the samples and elucidating their structures.
two techniques allows one to discriminate between two
possible structural isomers in a real forensic case sample
with high confidence.
Since no single technique can provide definitive structure Seized sample identification methods
elucidation, the Scientific Working Group for the Analysis Figure 1 shows the GC/MS chromatogram of a sample
of Seized Drugs1 (SWGDRUG) proposes a combination of seized from a clandestine lab. The unknown sample was a
analytical techniques be used depending on the nature of clear, highly volatile, and fragrant organic liquid, suspected
the samples and available techniques for analysis. Initial as a precursor compound used in the production of illicit
chemical identification of a seized sample includes a drugs. The data was acquired using a Thermo Scientific™
color test and FTIR analysis. Molecular weight information ISQ™ QD Single Quadrupole GC-MS system. The main
is provided by mass spectrometry. When combined component of the sample has a retention time of 4.38 min
with other spectroscopic techniques, including Raman (top trace) and the corresponding mass spectrum is shown
spectroscopy and NMR, a chemical structure can be at the bottom of Figure 1. The main component has a
elucidated. The power of GC/MS, FTIR and Raman is molecular mass of 202 g/mol. In addition, there are a series
highly dependent upon the databases of reference samples of fragmentation peaks at m/z 159, 129, 103, 73, and 57.
and related compounds. It is, therefore, an arduous The subsequent library search yielded a high probability
endeavor to maintain and update these databases in the match of the sample to 1,1-dibutoxy butane, with its
face of increasing numbers of novel structural analogues structure shown in Figure 2A.
that are appearing.
Figure 1
ISQ QD GC-MS
chromatogram of unknown
liquid (forensic case sample).
Figure 2
Chemical structures of two
isomers; (A) 1,1-Dibutoxybutane;
A B and (B) 1,1-Diisobutoxy-2-
methylpropane.
The sample was further analyzed by a picoSpin 80 1H The spectrum exhibits a distinct pattern of multiplicities
NMR spectrometer, an 82 MHz pulsed, Fourier transform characteristic to the isopropyl group: two strong overlapping
1
H NMR permanent magnet instrument, equipped with a doublets centered near 0.95 ppm, followed by a more
capillary cartridge probe. Since the sample was a clear complex series of weaker multiplets between 1 – 2 ppm. In
organic liquid, dilution in typical NMR solvents was not addition, two strong doublets emerge centered at 3.34 and
required. A small amount of the case sample was placed in 4.49 ppm, respectively. The pattern of doublets suggests
a vial, to which a few drops of tetramethylsilane (TMS) was the presence of a single neighboring proton (CH), and the
added to reference chemical shifts. The mixture was then shift to high frequency indicates that the carbon center
directly injected into the NMR spectrometer. The resulting is attached to an electron withdrawing group. The NMR
spectrum is shown in Figure 3. Note that the peaks spectrum also implies structural symmetry of the sample.
attributed to impurities and TMS are manually labeled gray, These suggested structural features, however, conflict with
whereas the unknown compound peaks are labeled blue. 1,1-dibutoxybutane suggested by the GC/MS library search.
Figure 3
Experimental 1H NMR 82 MHz spectrum of the seized sample. Data was acquired using a 90° pulse and 15 s recycle delay between pulses. The spectrum is an average of 5 co-added
scans and processed using the Mnova NMR software suite (Mestrelab Research Inc.).
Figure 4 shows the comparison between the predicted The key features of the NMR spectrum are assigned as
NMR spectrum of 1,1-dibutoxybutane by Mnova software follows:
(top) and the experimental 1H NMR spectrum of the seized
• An intense, overlapping set of doublets at 0.9 ppm
sample (bottom). There are vast differences between the
is due to the terminal methyl’s of the isopropyl group
two spectra, indicating the sample under analysis is not
(-CH(CH3)2);
1,1-dibutoxybutane.
• A broad, weak multiplet between 1 - 2 ppm is due to the
Upon further examination of the compound list from the methine (CH) coupling to the adjacent CH3;
GC/MS library search, a low probability match compound,
• The doublet centered at 3.3 ppm originates from the
1,1-diisobutoxy-2-methylpropane (Figure 2B), was identified
methylene groups (CH2) of the 2-methyl propyl moiety.
as a strong candidate. The structure possesses all key
The doublet arises from the coupling to the adjacent CH
elements suggested by the experimental NMR spectrum:
proton of the isopropyl fragment;
isopropyl groups, electron withdrawing O atoms, and high
symmetry. Figure 5 shows a good agreement between the • The doublet at 4.5 ppm belongs to the CH group. It shifts
predicted NMR spectrum of 1,1-diisobutoxy-2-methylpropane further due to their attachment to two electron withdrawing
(top) and the experimental NMR spectrum of the seized O atoms. The weaker signal compared to the doublet at
sample (bottom). 3.3 ppm is due to the presence of only one CH group.
Figure 4
Predicted NMR spectrum of 1,1-dibutoxybutane (top) and
experimental 1H NMR 82 MHz spectrum of the seized
sample (bottom). The predicted spectrum was generated
by using Mnova NMRPredict plugin. Field strength was
set at 82 MHz and linewidth value was set at 2.5 Hz.
Figure 5
Experimental 1H NMR 82 MHz spectrum of the seized
sample and the predicted NMR spectrum of 1,1-diisobutoxy-
2-methylpropane generated by Mnova software.
The GC/MS data in combination with the picoSpin 80
NMR spectrum allows for the positive identification of
the seized sample as 1,1-diisobutoxy-2-methylpropane.
The compound is a precursor chemical and not a variant
structural analogue of known illegal drugs.
Thermo Scientific™ picoSpin™ 80
NMR spectrometer and Thermo
Conclusions Scientific™ ISQ™ QD Single
In this note we showed how the picoSpin 80 NMR was Quadrupole GC/MS System
able to provide additional information by interpreting
and predicting the likely structure of a seized sample,
complementing those derived from GC/MS analysis. The
combination of the two techniques allows for a positive
identification of an unknown compound with high confidence.
The practices demonstrated in this note conform to the
recommendations by SWGDRUG that multiple techniques
are required to confirm seized sample identity.
References
SWGDRUG home page: http://www.swgdrug.org/
(accessed Jun 17, 2016).
For Research Use Only. Not for use in diagnostic procedures. © 2016 Thermo Fisher Scientific Inc. All rights
reserved. Mnova is a trademark of Mestralab Inc. All other trademarks are the property of Thermo Fisher Scientific
and its subsidiaries unless otherwise specified. AN52889_E 08/16M
APPLICATION NOTE
No. AN52907
Determination of polymer molecular weight and
composition using picoSpin NMR spectroscopy
Authors:
Katherine Paulsen and Daniel Frasco,
Thermo Fisher Scientific, Madison, WI
Key words
Abstract
Polymer molecular weight determination and copolymer by comparing the resonance signals from repeating
compositional analysis involve the integration of the units to those of end groups, thereby determining the
resonance signals from polymer repeating units and end molecular weight. A similar methodology can be adapted
groups, which are inherently broad. A Thermo Scientific™ for the compositional analysis of copolymers, so long as
picoSpin™ 80 NMR spectrometer is well suited for these the resonance signals from different monomers can be
analyses. In the example of poly (ethylene glycol) (PEG) clearly differentiated.
acetyl triarm, the number average molecular weight
was determined with great ease. In the case of the Due to poor molecular rotation and marginally different
compositional analysis of a commercial PEG-PPG-PEG chemical environments in which the repeating units are
block copolymer, Pluronic® L-35, the PEG/PPG ratio situated, resonance signals from polymer repeating units
determined by a picoSpin 80 NMR agrees well with the often coalesce as a broad peak, even using high-field
manufacturer’s product specification. Furthermore, the NMR spectrometers.3 However, in the molecular weight
analyses using an 82 MHz and a 300 MHz NMR yielded analysis and compositional analysis, collective signals
almost identical results. from polymer repeating units are used for calculation
and monomeric resolution is generally not required. To
Introduction that end, low-field NMR readily lends itself as a low-cost
The control of the molecular weight and molecular weight alternative to high-field instruments, with significant savings
distribution (MWD) is essential to obtain and improve on both instrument procurement and upkeep.
certain desired physical properties in a polymer product,
and is therefore of great importance in material science1. In this application note, molecular weight determination of
In addition to gel permeation chromatography (GPC) poly(ethylene glycol) (PEG) acetyl triarm and compositional
and matrix assisted laser desorption ionization mass analysis of a polyol using a picoSpin NMR spectrometer
spectrometry (MALDI MS)2, end group analysis using are presented. The results of the compositional analysis
nuclear magnetic resonance (NMR) spectroscopy has using an 82 MHz NMR were also compared with those
long been established as a viable analytical technique using a high-field, 300 MHz NMR spectrometer.
for determining the number average molecular weight
(Mn) of polymers. For polymers with a defined end group
structure, the number of repeating units can be deduced
Experimental Results and discussion
The structures of the polymers examined in this study are Number average molecular weight (Mn) determination.
shown in Figure 1. In 1H NMR spectroscopy, the area under each resonance
signal is proportional to the molar concentration of the
A B protons being analyzed. Number average molecular
weight (Mn) determination by end-group analysis using
1
H NMR therefore involves identifying and integrating
distinguishable proton signals originating from end-
groups and repeating units.
Figure 1: Chemical Structures of: A) poly(ethylene glycol) (PEG) acetyl triarm (PEG acetyl
triarm), and B) poly(ethylene glycol)-block-poly(propylene glycol)-block-poly(ethylene
glycol) (PEG-PPG-PEG).
known signal in the spectrum. MF =FWIJKLMNOJ JPFQRIM +FWMFS INOTUV + FWrepeating units =41+177+(44×7×3)=1142
integral of methyl protons 3.00
All spectra were acquired using the following acquisition Based on the end
x = group analysis using = an 82 MHz
# of methyl group protons 3
parameters: 90° excitation pulse, 750 ms acquisition time picoSpin NMR x = 1.0 relative moles of propylene glycol
spectrometer, the number average
and 5 second recycle delay. Spectral data were processed molecular weight of the PEG acetyl triarm is 1142 g/mol.
using the Mnova™ NMR analysis software with a standard 7.65 3.00
=
set of processing parameters including: Zero filling and 4y + 3 3
y = 1.16 relative moles of ethylene glycol
phase correction. Apodization was not used.
relative moles PEG×MWefg
weight % PEG = ×100%
relative moles PEG×MWefg + relative moles PPG×MWeeg
1.16×44
= ×100% = 46.8%
1.16×44 + 1.0×58
Compositional analysis of a PEG/PPG block copolymer.
Figure 3 shows the 1H NMR spectra of Pluronic L-35, with a normalized peak area of 3.0, and the CH- and
a PEG-PPG-PEG copolymer, using (a) an 82 MHz CH2-proton signals are located at δ ~ 3.3-3.7 ppm with
picoSpin NMR spectrometer; and (b) a 300 MHz NMR a normalized peak area of 7.77. The minor variances in
spectrometer. The two spectra are very similar. In the chemical shift between the two spectra are likely due
82 MHz spectrum, the signals at δ ~0.8 ppm with a to the difference in sample concentrations. Because of
normalized peak area of 3.0 are attributed to the methyl the higher sensitivity of a 300 MHz NMR instrument,
(-CH3) group of PPG. The CH- and CH2- proton signals the samples were used in a lower concentration; hence,
from both PPG and PEG are located at δ ~ 3.1-3.4 ppm, the difference in chemical shift. The 300 MHz spectrum
with a normalized peak area of 7.65. The contributions also offers greater resolution than the 82 MHz one. The
to these signals include 3 protons from PPG (one CH- increased resolution, however, should have little to no
and one CH2-) and 4 protons from PEG (2 CH2- groups). bearing on the results of the compositional analysis,
The 300 MHz 1H NMR spectrum is essentially the same: because the collective signals are integrated for the
the signal from the methyl group resonates at δ 1.1 ppm ensuing calculations.
B
A detailed compositional analysis using the 82 MHz NMR Conclusions
spectrum is outlined below.
peak area of repeating units and glycerol linkage
=
peak area of end groups
1
H NMR spectroscopy has been established as a powerful
# of protons in repeating units and glycerol linkage # of protons in end groups
tool for polymer characterization, including molecular
5.78 =.>>
= , n = 6.82 ≈ 7 repeating monomer units
1. peak area of repeating units and glycerol linkage
Determine the
12n + 5 5relative moles of PPG, denoted as x,
peak area of end groups
weight determination and copolymer compositional
MF =FWIJKLMNOJ JPFQRIM +FWMFS INOTUV + FWrepeating units=
=41+177+(44×7×3)=1142
using the signals at δ ~0.8.
# of protons in repeating units and glycerol linkage
peak area of repeating units and glycerol linkage # of protons in end groups
peak area of end groups
analysis. Both analyses involve the integration of the
=
5.78 =.>>
# of protons in repeating units and glycerol linkage
= # of protons in end groups
, n = 6.82 ≈ 7 repeating monomer units resonance signal from polymer repeating units, which are
12n + 5 5
integral of methyl protons 3.00
5.78
MF =FWIJKLMNOJ JPFQRIM
12n + 5 +FW
=.>>
=x = MFS INOTUV
, n = 6.82 =
≈ 7 repeating monomer units
+ FW repeating units =41+177+(44×7×3)=1142
5# of methyl group protons 3
inherently broad. The Thermo Fisher Scientific picoSpin 80
MF =FWIJKLMNOJ JPFQRIMx = 1.0 relative moles of propylene glycol
+FWMFS INOTUV + FWrepeating units =41+177+(44×7×3)=1142 NMR spectrometer is well suited for these investigations.
integral of methyl protons 3.00 In the example of poly(ethylene glycol) (PEG) acetyl triarm,
x = =
2. Determine the # of methyl group protons
relative moles3.00 of PEG,
integral of methyl protons
7.65 3denoted as y,
3.00 the number average molecular weight was determined with
x = = =
x = 1.0 relative moles of propylene glycol
# of methyl group protons 3
using the signals at δ4y + 3~3.1-3,4. 3
great ease. In the case of the compositional analysis of a
x = 1.0 relative moles of propylene glycol
y = 1.16 relative moles of ethylene glycol
commercial PEG-PPG-PEG block copolymer, Pluronic®
7.65 3.00
4y + 3
7.65
=
3
3.00
L-35, the PEG/PPG ratio determined by a picoSpin
=
y = 1.16 relative moles of ethylene glycol
4y + 3 3
relative moles PEG×MW
80 NMR agrees well with the manufacturer’s product
efg
weight % PEG = ×100%
y = 1.16 relative moles of ethylene glycol
relative moles PEG×MW efg + relative moles PPG×MWeeg specification. Furthermore, the analyses using an 82 MHz
3. Calculate
=
the weight
1.16×44 percentages of PEG and PPG.
×100% = 46.8%
and a 300 MHz NMR yielded almost identical results. The
1.16×44 + 1.0×58 relative moles PEG×MWefg
weight % PEG = ×100% Thermo Fisher Scientific picoSpin 80 NMR spectrometer
relative moles PEG×MW efg + relative moles PPG×MWeeg
relative moles PEG×MW
weight % PEG =
1.16×44 relative moles PPG×MW
efg
×100% has proven to be a low-cost alternative to high-field NMR
relative moles PEG×MW efg + relative moles PPG×MW
eeg eeg ×100%
weight % PPG =
= ×100% = 46.8%
1.16×44 relative moles PEG×MW
+ 1.0×58
1.16×44 efg + relative moles PPG×MWeeg spectrometers for polymer molecular weight determination
= ×100% = 46.8%
=
1.0×58
1.16×44 + 1.0×58
×100% = 53.2% and compositional analysis.
1.16×44 + 1.0×58 relative moles PPG×MWeeg
weight % PPG = ×100%
relative moles PEG×MW efg + relative moles PPG×MWeeg
relative moles PPG×MW
weight % PPG =
1.0×58
eeg
relative moles PEG×MWefg + relative moles PPG×MWeeg
×100% References
= ×100% = 53.2%
1.16×44 + 1.0×58
1.0×58
= ×100% = 53.2%
1.16×44 + 1.0×58 1. Higginbotham, C. L.; Izunobi, J.U. J. Chem. Ed. 2011,
88, 1098-1104
Based on the 82 MHz NMR spectrum, Pluronic L-35
copolymer contains 46.8 % of PEG and 53.2 % of PPG. 2. Skoog, D. A.; Holler, J. F.; Crouch, S. R. Principles
The results using the 300 MHz spectrum are essentially the of Instrumental Analysis, 6th ed.; Brooks Cole: Pacific
same: 47.4 % of ethylene glycol and 52.6 % of propylene Grove, 2006.
glycol. Both results are in good agreement with the product
specification4, where PEG/PPG (w/w) = 1. 3. Bovey, F. A.; Mirua, P. A. NMR of Polymers, 1st ed.;
Academic Press: San Diego, 1996.
For Research Use Only. Not for use in diagnostic procedures. © 2016 Thermo Fisher Scientific Inc. All rights
reserved. Mnova is a trademark of Mestralab Inc. All other trademarks are the property of Thermo Fisher Scientific
and its subsidiaries unless otherwise specified. AN52907_E_11/16M
APPLICATION NOTE
No. AN52948
Qualitative and quantitative analysis of
the polymerization of PS-b-PtBA block
copolymer using picoSpin 80 NMR
Authors:
Katherine Paulsen and Daniel Frasco,
Thermo Fisher Scientific, Madison, WI, USA
Yufeng Zhu, Department of
Macromolecular Science and Engineering,
Case Western Reserve University,
Cleveland, OH, USA
Key words
picoSpin 80, NMR, polymerization, copolymer, number of
repeating units, polystyrene
Benefit
The Thermo Scientific™ picoSpin™ 80 NMR spectrometer
offers near real-time analytical capability for reaction
monitoring due to its unique capillary injection design Introduction
and compact footprint. The picoSpin 80 NMR has also Nuclear magnetic resonance (NMR) spectroscopy has
proven to be a low-cost alternative to high-field NMR proven to be an invaluable analytical tool for reaction
spectrometers to quantitatively assess polymerization and monitoring and optimization by elucidating molecular
to determine copolymer structure. structure, studying reaction kinetics, monitoring reaction
progress and gauging product purity1. For polymers
Abstract however, resonance signals from repeating units often
The use of the low-field picoSpin 80 NMR spectrometer coalesce as broad peaks, even with high-field NMR
offers a near real-time analytical capability for reaction spectrometers. This is largely due to poor molecular
monitoring, largely due to its unique capillary injection design rotation and repeating units being situated in marginally
and compact footprint. Despite the broad resonance signals different chemical environments. Low-field NMR such as
typical for polymers, by carefully selecting the regions where the picoSpin 80 NMR spectrometer lends itself as a low-
signal changes manifest the involved chemical transformation, cost alternative to high-field instruments for polymerization
the integrated peak area provides valuable insight to the monitoring, resulting in significant cost savings on both
reagent conversion as well as the final product structure. instrument procurement and upkeep.
Presented herein is a case study using a picoSpin 80 set of processing parameters including zero filling and
NMR spectrometer to monitor the polymerization of t-butyl phase correction. Apodiation was not used.
acrylate with a polystyrene reagent. The reaction scheme
is shown in Figure 1. By carefully identifying and integrating Results and discussion
the resonance signals associated with the monomer, the A qualitative look at the polymerization
progress of the polymerization was successfully assessed Figure 1 depicts a narrowly dispersed polystyrene-based
both qualitatively and quantitatively. The number of t-butyl reversible addition-fragmentation chain transfer (RAFT)
acrylate units incorporated into the final copolymer product reagent reacting with t-butyl acrylate monomer to form a
was also deduced. PS-b-PtBA block copolymer2-3. Figure 2 shows the NMR
spectrum of the initial reaction mixture containing t-butyl
acrylate (tBA) and polystyrene reagent (PS) in anisole. The
signals between δ 1.4–2.8 ppm originate from the 3 protons
of the CH2- and CH- groups of the PS repeating unit. These
overlap with the singlet from the 9 t-butyl methyl protons
Figure 1: Polymerization of t-butyl acrylate with a polystyrene reagent to form PS-b-PtBA. of tBA (δ 1.7 ppm). The singlet at δ 3.7 ppm is ascribed to
the methyl group of anisole. The signals between δ 5.5 –
Experimental 6.7 ppm correspond to the 3 vinyl protons of tBA. Finally,
Monomer t-butyl acrylate (tBA, 0.150 g, 1.170 mmol), the multiplet in the region δ 6.8 – 7.6 ppm arises from the
polystyrene reagent (PS, 0.230 g, 0.120 mmol), 2,2’-azobis aromatic rings from both anisole and PS.
(2-methylpropionitrile) radical initiator (AIBN, 0.001 g, 0.006
mmol), and anisole (1.156 mL) were added together into a
10 mL flask. The number average molecular weight of PS
is 1910.76 g/mol, determined by the end-group analysis
using a 500 MHz NMR and confirmed by GPC. The
average number of repeating units in PS is 16. This resulted
in a molar ratio of 9.75/1 for tBA/PS. The mixture was
degassed by three freeze-pump-thaw cycles. The mixture
was then stirred at 75 °C for 12 h and the polymerization
process was observed by 1H NMR spectroscopy.
Figure 5: 82 MHz 1H NMR spectrum of the polymerization process after 12 hours at 75 °C.
Figure 4: Schematic showing the conversion of vinyl protons to protons in the polymer
backbone. The numbers denote the projected chemical shifts of the involved protons by
Based on the calculations below, it was determined that
1
H NMR. 34% of tBA was converted into the block copolymer,
corresponding to an average of 34% of 9.75 = 3 tBA units
It is worth pointing out that the compact footprint of the in every copolymer molecule.
picoSpin 80 NMR spectrometer allows the instrument
to be set up in the laboratory where the chemistry takes
place. Furthermore, the unique capillary injection system
of the picoSpin minimizes sample workup, offering a
near real-time analytical capability critical for reaction
monitoring and optimization.
2. Li, B.; Shi, Y.; Zhu, W.; Fu, Z.; Yang, W. Polym. J. 2006,
38, 387-394
For Research Use Only. Not for use in diagnostic procedures. © 2017 Thermo Fisher Scientific Inc. All rights
reserved. All trademarks are the property of Thermo Fisher Scientific and its subsidiaries unless otherwise specified.
AN52948_E 03/17M
APPLICATION NOTE
No. AN52976
Analysis of acetanilide herbicides and their
rotational isomers by picoSpin 80 NMR
Author:
Daniel Frasco, Thermo Fisher Scientific,
Madison, WI, USA
Key words: NMR, agrochemical,
pesticide, herbicide, isomers
Application benefit
Thermo Scientific™ picoSpin™ 80 Series II NMR is a
powerful analytical tool to qualitatively and quantitatively Due to the widespread use of these herbicides, it is
analyze acetanilides and their isomers. Compared to the important to understand the metabolic pathways of these
traditional methods using chromatography and high-field molecules and identify their transformation products. Special
NMR, the NMR method using picoSpin 80 is cost-effective, attention has to be paid to stereoisomers, as they could
time-efficient and straightforward. have vastly different reactivity and degradation pathways.3-4
For acetanilides, amide cis-trans isomerization has been well
Abstract established where the rotation of the amide bond from the
The picoSpin 80 Series II NMR spectrometer can be used as a substituted aromatic ring is hindered.3-6
low-cost alternative to a high-field instrument for the qualitative
and quantitative analysis of acetanilide herbicides. The Quantitative determination of the isomeric ratio in a mixture
picoSpin offers adequate resolution to differentiate between presents a particular analytical challenge. Traditionally, a
acetanilide isomers and a facile means to determine the number of techniques including Gas Chromatography-Mass
isomeric ratio by simply integrating the corresponding signals. Spectrometry (GC-MS), Capillary Zone Electrophoresis (CZE),
High-Performance Liquid Chromatography (HPLC) and
Introduction high-field Nuclear Magnetic Resonance (NMR), have been
Acetanilides are a class of used to analyze the acetanilide isomers.4-5 Chromatographic
preemergent herbicides that inhibit techniques including GC-MS, HPLC and CZE often require
the growth of unwanted grasses time-consuming method development. The chromatography
and weeds associated with the runs are generally on the order of 10-15 min, often with
production of agricultural crops like Figure 1: General Structure of non-baseline resolution, giving rise to the uncertainty in
corn, soybeans, cotton and peanuts.1 Acetanilide Herbicides isomeric ratio. High-field NMR has proven to be a powerful
The general structure of acetanilide tool for the analysis of acetanilide isomers by integrating their
herbicides is shown in Figure 1. signals. However, the associated intensive capital and space
investment often prohibits their use in smaller facilities or in
Acetanilides represent some of the most widely used factories where pesticide production is taking place.
pesticides for both crop and non-crop use. In 2007,
metolachlor and acetochlor, two representative acetanilides,
were the 4th and 5th most used pesticides in the United
States, with an approximated use of around 30 million
pounds each.2
This work demonstrates the qualitative and quantitative
analyses of a group of acetanilides with varying structural
complexity using an inexpensive and portable picoSpin 80
Series II NMR spectrometer. NMR spectra were obtained
in as little as a few seconds compared to the longer
analysis times of other techniques. Additionally, a facile
means to determine the isomeric ratio is also presented.
This methodology was applied to investigate the impact of
solvent polarity on the ratio of cis-trans rotational isomers
of the acetanilide.
Experimental
A picoSpin 80 Series II NMR spectrometer was used
to acquire the spectra of acetanilide herbicides. The Figure 2: 1H NMR spectrum of a 1.8M solution of propachlor in DMSO-d6 acquired in 5
spectrometer is an 82 MHz, pulsed, Fourier transform minutes using 32 scans. The residual DMSO signal and water present in the deuterated
solvent are identified.
1
H NMR. The instrument contains a 2 Tesla temperature
controlled permanent magnet and is fitted with a 40 in Figure 3 clearly indicates that the resolution of the picoSpin
microliter capillary cartridge used for sample introduction 80 Series II is sufficient to resolve the signals corresponding
into the spectrometer. to each isomer. In the case of alachlor, ethyl groups are
located at the 2 and 6 position of the aromatic ring. The
All samples were purchased from Sigma Aldrich and were orientation of the carbonyl group away from the aromatic ring
used as received. Both solid and liquid samples were corresponds to the trans- isomer, and the one aligning toward
analyzed in deuterated solvent and injected into the capillary the aromatic ring corresponds to the cis- isomer. In Figure 3,
cartridge using 1 mL slip-tip polypropylene syringes and 22 the three methoxy protons from the two isomers can be seen
gauge blunt-tipped needles. Back-to-back sample injection as two singlets at δ 3.3 and 3.4 ppm. The two methylene
and data acquisition was separated by a solvent/air/solvent/ protons next to the carbonyl show a significant difference in
air flush to thoroughly clean the capillary. chemical shift due to their change in proximity to the aromatic
ring in the two isomers. They are present as singlets at δ 3.8
All spectra were acquired using the following acquisition ppm (trans-) and δ 4.7 ppm (cis-). The two methylene protons
parameters: Between 1 and 64 scans, 90° excitation pulse, located between the amide nitrogen and the methoxy oxygen
750 ms acquisition time and 8 second recycle delay. Spectral resonate at δ 4.9 ppm for both isomers, although a small
data was processed using the Mnova™ NMR analysis shoulder ascribed to the minor cis- isomer can be observed,
software with a standard set of processing parameters. indicating a subtle difference caused by the isomerism.
Conclusions
The picoSpin 80 Series II NMR spectrometer can be used
as a low-cost alternative to a high-field instrument for the
qualitative and quantitative analysis of acetanilide herbicides.
Some acetanilide molecules are present as rotational
isomers due to hindered rotation of the amide bond.
The picoSpin offers adequate resolution to differentiate
the isomers and a facile means to determine the isomeric
ratio by simply integrating the corresponding signals. The
results clearly suggest that solvent polarity has a profound
impact on the isomeric distribution in solution, and great
caution should therefore be exercised when analyzing and
interpreting the isomeric ratio.
References
1. Morton, M. D.; Walters, F. H.; Aga, D. S.; Thurman, E. M.;
picoSpin 80 Series II NMR spectrometer
Larive, C. K. J. Agric. Food Chem. 1997, 45, 1240-1243
For Research Use Only. Not for use in diagnostic procedures. © 2017 Thermo Fisher Scientific Inc. All rights
reserved. All trademarks are the property of Thermo Fisher Scientific and its subsidiaries unless otherwise specified.
AN52976 0517
APPLICATION NOTE AN52979
Table 1: Average hydrogen content and % relative standard deviation Table 2: Accuracy of picoSpin 80 NMR spectrometer for determining
of five runs per reference sample determined using picoSpin 80 NMR. hydrogen content using five runs per reference sample.
Figure 3: Full 19F NMR spectrum of perfluoro-n-butyl ethylene. Inset: Figure 4: Full 1H NMR spectrum of 2H,3H-decafluoropentane (C5H2F10;
Multiplet analysis of the splitting pattern of individual peaks, with a neat) with TMS added as a chemical shift reference. Multiplet analysis
J-coupling tree and coupling constants overlaid on the spectrum. of the splitting pattern reveals a dddd class; a J-coupling tree and
coupling constants are overlaid on the spectrum.
In the 19F spectrum (Figure 5), primary and secondary Conclusions
alkyl fluorides experience vastly different shielding along Fluorine-19 NMR compliments 1H and 13C NMR in
the carbon backbone, spanning a range from -220 ppm structure determination. The broader chemical shift range
to -75 ppm, allowing for the analysis of individual fluorine in 19F NMR helps resolve individual fluorine containing
groups separately. Confirmation of the strong geminal 2JFH functional groups, while the often-large variable magnitude
coupling observed in the 1H spectrum is in the multiplet of 19F-19F and 1H-19F coupling provides additional insight
analysis of the splitting pattern of fluorine atoms at into structural effects. First-order coupling and highly
positions C2 and C3 (inset for peaks 2 and 3). resolved resonance lines also simplify analysis. Overall,
19
F NMR adds a new dimension in the analysis portfolio of
fluorine containing compounds.
Bulk PEs are manufactured as pellets (resins, granules), and later converted to
other forms (such as films and pipes) using extrusion or molding processes.
They are also made into multilayer films for a wide range of industrial
applications like food and consumer product packaging. The density of bulk
PE pellets and single-layer PE films can be measured and classified with
relative ease using several standard techniques: ISO 1183-1/ASTM D792
(immersion method),3 ISO 1183-2/ASTM D1505 (density gradient method),4
and ASTM D4883 (ultrasound method).5 However, all these techniques
require the PE in its “pure” form, which can be challenging in the case of PE
in multilayer films. Extensive sample preparations (microtoming, separation
of layers by dissolving in solvents) are often required6 to isolate the PE layer
before analysis, which can be labor-intensive and time-consuming.
Raman spectroscopy is sensitive to changes in the A 532 nm laser was used with a 2 mW laser power at the
molecular structure level of PE, such as the degree of sample. A 10x objective and a 50 μm slit aperture were used
crystallinity, which is the key determining factor of PE to obtain more representative spectra from the samples.
density.7,8 More importantly, the confocal capability of Total acquisition time for each spectrum was 30 seconds
Raman microscopy allows for facile in situ analysis of (3 second exposure x 10 exposures). Thermo Scientific™
individual PE layers in multilayer films without the need to OMNIC™ software was used for operation of the DXR2
isolate the PE layer. To our best knowledge, PE density Raman Microscope, and collection of Raman spectra;
measurement using Raman has been limited to PE Thermo Scientific™ TQ Analyst™ software was used for
pellets.7,8 In this work, we want to systematically explore chemometric analysis of the Raman data.
the feasibility of using confocal Raman microscopy for PE
film density analysis, both qualitatively and quantitatively. Results and discussion
We demonstrate that Raman microscopy in combination Raman spectra
with the discriminant analysis method can be successfully Representative Raman spectra of HDPE and LDPE
applied to distinguish HDPE and LDPE in both pellet samples, in both pellet and film forms, are shown in
and film forms. In a subsequent application note, we will Figure 1. There are noticeable differences between HDPE
detail the quantitative determination of PE density using a and LDPE spectra, for both pellets and films. In the CH2
confocal Raman microscope. bending and the CH2 twisting region, the intensity of
the CH2 bending mode at 1416 cm-1 (relative to the CH2
Experimental bending mode at 1440 cm-1) is higher for HDPE than for
Sample description LDPE. This observation agrees with the previous reports
A total of 16 PE samples (10 pellets and 6 films) with that the 1416 cm-1 and 1440 cm-1 peaks are indicators of
known densities were used for the classification studies. crystalline and amorphous PE phases, respectively.7-10
All samples were used as received. The higher the degree of crystallinity, the higher the
density. The differences between HDPE and LDPE
Method conditions are also pronounced in the C-H stretching region. The
A Thermo Scientific™ DXR2™ Raman Microscope was intensity of the symmetric CH2 stretching mode at 2848
used for the collection of Raman data. For each type/class cm-1 (relative to the asymmetric CH2 stretching mode at
of the pellet samples, Raman spectra were collected from 2882 cm-1) appears to be higher for LDPE compared to
3 different pellets and averaged. For each film sample, HDPE. Since the C-H stretching (2825-2970 cm-1) and the
Raman spectra were collected from 3-4 locations across CH2 bending regions (1398-1470 cm-1) are sensitive to the
the surface of the sample. An averaged spectrum was density of PE, these regions were selected for subsequent
then used for final analysis. discriminant analysis.
Figure 1: Representative Raman spectra of HDPE and LDPE pellets and films. (A) Full spectral range. (B) C-H stretching region. (C) CH2 bending and
CH2 twisting region.
Data processing
The raw Raman spectra were processed using Norris Norris derivative is effective in removing background drift
2nd derivative, and the resulting spectra were further in Raman spectra caused by fluorescence, whereas SNV
processed by standard normal variate (SNV). Examples of is effective in compensating such variations as sample
the data processing are shown in Figure 2. surfaces and laser penetration depths.11-12
Figure 2: Norris 2nd Derivative and SNV processed sample spectra. (A) Full spectral range. (B) C-H stretching region. (C) CH2 bending region. HDPE
spectra are in red in both B and C Plots.
Calibration results
PE sample Actual class Usage Calculated class Distance to HDPE Distance to LDPE
1 HDPE Calibration HDPE 0.89 4.23
2 HDPE Calibration HDPE 0.77 4.97
5 HDPE
Pellets
Table 2: Impact of number of PCs on variance coverage Figure 3: 3-D cluster plot of HDPE and LDPE samples. The • are the
calibration samples, and the + are the validation samples.
Figure 5: Classification of a PE sample by using its Raman spectrum and the discriminant method created by the TQ Analyst software.
Conclusions
In this application note, we have successfully
demonstrated the use of a Thermo Scientific DXR2 Raman
Microscope, in combination with the TQ Analyst software,
to classify polyethylene’s of different density classes in
both pellet and film forms. Raman spectroscopy is non-
destructive and requires minimal sample preparation.
The classification method was created solely based on
the Raman spectral features of LDPE and HDPE and
was indifferent to the sample forms. Once the method is
established, PE samples, pellets or films, can be correctly
classified within minutes. Moreover, this work expands
the scope of the previously reported study on PE pellets
to include PE films, which broadens its applicability in the
Thermo Scientific DXR2 Raman Microscope
plastic/polymer industry as well as many downstream
industries. The described methodology should be
applicable for in situ classification of thin PE layer(s) in
multilayer films.
References
1. Piringer O.G. and Baner A.L., ed. Plastic Packaging: Interactions with Food and Pharmaceuticals. 2nd ed. Weinheim:
Wiley-VCH; 2008.
2. Polyethylene, The Essential Chemical Industry - online,
http://www.essentialchemicalindustry.org/polymers/polyethene.html, retrieved on 9/10/2017.
3. (a) ISO 1183-1:2012, Plastics − Methods for determining the density of non-cellular plastics − Part 1: Immersion
method, liquid pyknometer method and titration method; ANSI New York, NY. (b) ASTM D792-13, Standard test
methods for density and specific gravity (relative density) of plastics by displacement; ASTM International, West
Conshohocken, PA, 2013.
4. (a) ISO 1183-2:2004, Plastics — Methods for determining the density of non-cellular plastics — Part 2: Density
gradient column method. (b)ASTM D1505-10, Standard test method for density of plastics by the density-gradient
technique; ASTM International, West Conshohocken, PA, 2010.
5. ASTM D4883-08, Standard test method for density of polyethylene by the ultrasound technique; ASTM International,
West Conshohocken, PA, 2008.
6. Mieth A., Hoekstra E., and Simoneau C. Guidance for the identification of polymers in multilayer films used in food
contact materials: User guide of selected practices to determine the nature of layers, EUR 27816 EN, 2016; doi:
10.2788/10593.
7. Strobl G.R. and Hagedorn W. Raman spectroscopic method for determining the crystallinity of polyethylene, J.
Polym. Sci. B Polym. Phys., 1978, 16, 1181-1193.
8. Sato H., Shimoyama M., Kamiya T., Amari T., Sasic S., Ninomiya T., Siesler H.W., and Ozaki Y. Raman spectra of
high-density, low-density, and linear low-density polyethylene pellets and prediction of their physical properties by
multivariate data analysis, J. Appl. Polym. Sci., 2002, 86, 443–448.
9. Migler K.B., Kotula A.P., Hight Walker A.R. Trans-rich structures in early stage crystallization of polyethylene,
Macromolecules, 2015, 48, 4555−4561.
10. Kida T., Hiejima Y., and Nitta K-H. Raman spectroscopic study of high-density polyethylene during tensile
deformation, Int. J. Exp. Spectroscopic Tech., 2016, 1:001.
11. Simone, E., Saleemi A.N., and Nagy Z.K. Application of quantitative Raman spectroscopy for the monitoring of
polymorphic transformation in crystallization processes using a good calibration practice procedure, Chem. Eng.
Res. Des. 2014, 92, 594–611.
12. Huang J., Romero-Torres S., and Moshgbar M. Practical considerations in data pre-treatment for NIR and Raman
spectroscopy, Am. Pharm. Rev. 2010, 13 (6).
13. Thermo Scientific Product Overview. TQ Analyst Software Chemometric Algorithms, 2009.
Figure 4: Raman spectra of a seized street drug sample containing cocaine and β-D-lactose.
Application benefits The analytical tools for laminate analysis include optical microscopy, differential
With a properly configured scanning calorimetry (DSC), Fourier Transform infrared spectroscopy (FT-IR),
Thermo Scientific™ DXR™2 Raman and Raman spectroscopy. Confocal Raman microscopy, in particular, offers
microscope, it was possible to resolve many advantages. Raman spectroscopy is sensitive to both chemical and
a thin polymer layer with a thickness physical properties. Its unique selection rules generate a molecular fingerprint
(≈0.4 μm) that’s very close to the that is well suited to material identification. Confocal Raman microscopy often
theoretical spatial resolution limit. utilizes short-wavelength visible and NIR (400 – 785 nm) lasers for sample
excitation, which renders increased sensitivity because Raman signal intensity
Thermo Fisher Scientific is inversely proportional to the 4th power of the laser wavelength. Furthermore,
solutions the spatial resolution is also inversely proportional to the laser wavelength. The
• DXR2 Raman microscope shorter the wavelength, the higher the resolution.1,2
A B
C D
Figure 1. (A) Video image of microfilm A showing the region where Raman line mapping was performed (red line); (B)
Raman line map of microfilm A; (C) representative Raman spectra of each layer. PE = polyethylene, PP = polypropylene,
PVA = poly vinyl alcohol; and (D) Raman correlation profile obtained using a PVA reference spectrum.
Figure 2A-2C shows the 3-D Raman area images of correlation profiling using a PVA spectrum from the map
microfilm B. PE, PP and PVA spectra were carefully as the reference, and the result is shown in Figure 2D.
selected from the map as the references to construct The sharp peak at ~ 14.5 μm corresponds to the layer 4,
the Raman area images with correlation profiling. The with an estimated thickness of ~ 0.4 μm based on the
red color indicates a high correlation with the reference FWHM of the profile peak.3 It’s important to note that
material and the blue color indicates low correlation. the methodology described above relies heavily on the
Similar to microfilm A, seven layers are clearly identified: quality of the spectra as well as the spectral differences
the layers 1, 3, 5 and 7 are PE, the layers 2 and 6 are PP, between adjacent layers. The layer thickness derived from
and the layer 4 at the center is PVA. In order to estimate the correlation profile should therefore be considered an
the layer 4 thickness, a line map was first extracted estimate. Precise measurement of layer thickness requires
from the area map, delineated by the white, dashed such technique as scanning electron microscopy (SEM).
line in Figure 2C. The extracted line was then subject to
A B C
Figure 2. 3-D Raman correlation images for microfilm B. (A)-(C): Raman correlation images for PE, PP, and PVA, respectively. (D)
Raman correlation profile obtained using a PVA reference spectrum on a line extracted from the area map (white dashed line on C).
Diffraction-limited resolution in optical microscopy is often sample that is optically denser than air, the axial resolution
empirically assessed by the Rayleigh criterion shown is usually 4-6 times lower the lateral one. For a 532 nm
below, where d denotes the Rayleigh criterion, λ is the laser and a 100× objective (N.A. = 0.90), assuming the
laser wavelength, η is the refractive index of the immersion/ refraction index 1.5 for the laminates, the theoretical lateral
mounting media, and N.A. is the microscope objective spatial resolution and axial resolution are approximately
numerical aperture. 0.4 μm and 2 μm, respectively. However, many factors
such as scattering of the laser/Raman photons and
interaction with interfaces in the sample can reduce this
resolution. Another important consideration in instrument
configuration is the pinhole size. The confocal pinhole acts
as a spatial filter, by allowing the Raman spectrometer to
look into a smaller spatial domain than with a conventional
configuration without the pinhole, attenuating the out-
As indicated in the equations, both lateral and axial Rayleigh of-focus regions of the sample.2 In the current study, the
criteria are directly proportional to the wavelength, but combination of a 532 nm laser, a 100× objective with an
inversely proportional to the objective numerical aperture. N.A. of 0.9, and a 25 μm confocal pinhole enabled the
In addition, the axial Rayleigh criterion is proportional to the resolution of a 0.4 μm-thick PVA layer, approaching the
refractive index of the material under investigation. For any theoretical limit of the lateral spatial resolution.
Conclusions
In this communication, we have demonstrated the
analysis of polymer laminates using Raman line and area
mapping. In both cases, seven layers were identified.
With a properly configured Raman microscope, it was
possible to resolve a thin polymer layer with a thickness
(≈0.4 μm) that’s very close to the theoretical spatial
resolution limit.
References
1. Guillory P., Deschaines T. Henson P, Materials Today,
2009, 12(4), 38-39.
Experimental
The cross section of an OTC bilayer tablet containing a blend of excipients and
an active pharmaceutical ingredient (API) was examined in this study. Spectra
were measured at 8 cm-1 resolution and 16 scans at each measurement point
using a Nicolet iS50 FTIR spectrometer, configured with a CaF2 beamsplitter,
InGaAs detector and iS50 Raman accessory (Figure 1). The laser wavelength
was 1064 nm with a spot size of approximately 50 microns and was applied
at a power of 500 mW. The area mapped was approximately 2 x 3 mm at 45
microns steps (Figure 2). Analysis of spectral maps was carried out using the
OMNIC Atlμs and OMNIC Specta software.
Position (micrometers)= -143 µm; -15745 µm; Number: 185
Composite for Match 1
+
a-Lactose Monohydrate
Beta-Cyclodextrin Hydrate
Figure 6: Top spectral plots show the spectra of the four identified components calculated by PCA: the experimentally
measured Raman spectra (blue), the PCA components (red), and library spectra (magenta) when available. Maps show
the distribution of each component across the measured area.
No. AN52910
Characterizing amber and its imitations with
dispersive Raman spectroscopy
Author:
Alex Rzhevskii, Thermo Fisher Scientific,
Madison, WI
Key words
Introduction
Amber is a tree resin that has been fossilized over
millions of years through polymerization, cross-linking
of the long carbon chain organics, and expulsion of
volatiles within the resin. Amber has a wide variety of
applications, ranging from its traditional use in jewelry,
art and craft objects, and folk medicine, to modern
applications in skin care products, perfumes, and While FT-Raman spectroscopy, in the near-infrared (1064
blood conservation devices. Authentic amber is a highly nm) region, has been used successfully to analyze amber1-3
demanded commodity and its high value has fostered an because it minimizes the substantial florescence originating
extensive imitation market that abounds with both natural from the amber resin, dispersive Raman spectroscopy
and synthetic resins. Copal is an example of one of the has important advantages over FT-Raman. The use of
natural resins that is used as an amber imitation. It is a lasers of shorter wavelength (e.g., 780 nm), especially,
less mature tree resin product that has not gone through in Raman microscopy allows for sample analysis on a
the entire fossilization process and, as a result, it is less sub-micrometer scale and results in significantly faster
expensive than amber. spectral acquisition, because of the high laser power
density developed in the small focused laser spots coupled
Distinguishing authentic amber from its imitations with more effective Raman scattering that is inversely
is important for scientific, commercial and personal proportional to the 4th power of the laser wavelength4.
reasons. There are several simple chemical and physical Furthermore, confocal Raman microscopy can provide
tests that can be used to authenticate amber. These useful information about the chemical composition of the
tests include solubility in organic solvents, floating in relict biomaterials and inorganic inclusions5.
salt water for testing specific gravity, assessing melting
points, and even evaluating hardness using scratch In this note, the spectra of amber and some of its
tests. The main limitation of those tests arises from common imitations obtained using a dispersive Raman
the possible compromise in specimen integrity, which spectrometer are compared and discussed.
can be prohibitive when dealing with items of artistic,
archeological or historical significance. To that end,
Raman spectroscopy offers an attractive solution to the
identification and characterization of amber, largely due to
its non-destructive sample handling.
Experimental Discrimination between amber and copal is slightly more
Samples of authentic and altered amber and its common involved because of their chemical similarities. Figure 2
imitations were used in this study, including: 1) Baltic shows the Raman spectra of amber and two copal samples
amber from a contemporary bead necklace, 2) a metal- of different origins. The most conspicuous features in
incrusted ear-ring made of, supposedly, a synthetic resin, all spectra are the bands at 1646 cm−1 and 1450 cm−1,
3) a pipe bit made of melted and pressed Baltic amber; resulting from the exomethylene stretching ν(C=C), and
4) West African copal, 5) Indonesian copal, 6) a sandarac the scissoring type deformation δ(CH2) vibration in resin
from Australia, 7) a pine tar, 8) a mastic resin. structure, respectively. It has been shown [2-4] that the
intensity ratio of the bands at 1646 cm−1 and 1450 cm−1
Raman spectra were acquired with a Thermo Scientific (/1646 ∕/1450) is a good indicator of the maturity of fossilized
DXR™ Raman microscope using 780 nm laser excitation resins. In general, the ratio is >1 for immature resins such
and 14 mW of power at sample. The typical exposure as copal, and <1 for mature resins such as amber. Our
time was set to 5 minutes to maximize signal-to-noise experiment results conform to that criterion. In Figure 2,
ratio. The real-time autofluorescence correction was the peak height ratio /1646 ∕/1450 is about 0.8 for amber (1) and
enabled to minimize fluorescence effects in the spectra. 2.5 and 1.1. for copal (2) and (3), respectively. In addition,
No further smoothing of spectral data was applied. there are noticeable differences between amber and copal
in the spectral region below 1400 cm-1, which is dominated
Results and discussion by a series of less intensive peaks, especially, in the 800-
Distinguishing amber from its synthetic forgeries by 500 cm-1 range. It was found that thermal and mechanical
Raman spectroscopy is rather straightforward. Figure 1 alteration of the amber leads to noticeable spectral changes
shows the “fingerprint” region of the Raman spectra of in the region of 1800 - 1400 cm-1, as shown in Figure 3.
an authentic Baltic amber and a synthetic resin, as well
as a reference spectrum of bisphenol-A. The spectrum
of the synthetic resin (Figure 1.2) bears little similarity
with that of amber (Figure 1.1), but has features that can
be unambiguously attributed to the bisphenol-A, the
component of an epoxy resin. Thus, in the “fingerprint”
spectral region, amber can be readily distinguished
from synthetic plastics due to a significant difference
in chemical composition and the identification of the
synthetic plastics can be conducted using a search
against extensive Raman spectral libraries of polymers.
Figure 2: Raman spectra of (1) melted and pressed Baltic amber from a pipe bit; (2)
Indonesian copal; and (3) West African copal. The spectra are normalized to the most
intense spectral bands.
Figure 1: Raman spectra of: (1) authentic Baltic amber from a contemporary bead
necklace; (2) synthetic resin from an ear-ring; and (3) bisphenol-A from the Nicolet
Standard Collection of Raman Spectra. The spectra are normalized to the most intense
spectral bands.
Figure 3: The Raman spectra of Baltic amber from bead necklace (red) and melted and
pressed amber from a pipe bit (brown).
The band at 1646 cm-1 becomes broader, and the intensity The spectral characteristics of amber are demonstrated to
of the shoulder at about 1615 cm-1 associated with the be vastly different from, for example, epoxy resin often used
stretching of aromatic C=C increases in the spectrum of for amber imitation. The fluorescence typical for amber,
the processed amber. In [6], these changes have been copal and other natural resins are effectively suppressed
observed and attributed to the increase of the aromaticity by using a 780-nm excitation, or it can be removed with
as the result of the thermal impact to the material. a real-time automated fluorescence correction procedure
available in Thermo Scientific™ OMNIC™ software. The
Raman spectroscopy can be used to differentiate amber use of dispersive Raman instrumentation along with visible
from other types of natural resins that look similar to amber. lasers enabled shorter sampling times in comparison with
Figure 4 shows the Raman spectra of three different resins. FT-Raman, where spectral acquisition times as long as
In the spectrum of mastic resin (1), the ratio /1646 ∕/1450 is <1 hours are often required5.
indicating its similarity to amber. However, the band at 1646
cm-1 in the spectrum of mastic resin exhibits a prominent References
shoulder at 1710 cm-1 less noticeable for amber (Figure 2.1)
that is ascribed to the stretching mode ν(C=O). For the pine 1. Beck C.W. Spectroscopic investigation of amber,
tar (2) and sandarac from Australia (3), the ratio /1646 ∕/1450 Applied Spectroscopy Reviews, Vol. 22, 57–110, 1986
is >1 and ≈1, respectively, typical for natural resins. In the
spectrum of pine tar (2), there is also a salient side peak at 2. Edwards H.G.M., Farwell D.W. Fourier transform-
about 1615 cm-1. Raman spectroscopy of amber, Spectrochimica Acta,
Part A, Vol. 52, 1119-1125, 1996.
For Research Use Only. Not for use in diagnostic procedures. © 2016 Thermo Fisher Scientific Inc. All rights
reserved. All trademarks are the property of Thermo Fisher Scientific and its subsidiaries unless otherwise specified.
AN52910_E 12/16M
APPLICATION NOTE AN52984
Experimental
A transdermal nicotine patch sample was mounted onto a gold-coated
microscope slide with the backing layer facing the microscope objective, and
the release liner at the bottom. Raman confocal line depth profiling and area
depth profiling were performed with a DXR2 Raman microscope (Figure 1)
using 532 nm laser, 5 mW laser power at the sample, 50X objective, 25 µm
Figure 1: DXR2 Raman microscope
confocal pinhole aperture, and with auto exposure (S/N = 200). For the line
depth profiling (Z-profiling), a depth of 220 μm was probed by using a 5 μm
step size (containing 45 points or spectra). For the area depth profiling, a
vertical X-Z area of 120 μm (X) by 245 μm (Z) was probed by using a 5 μm step
size along the Z-axis and 20 μm along the X-axis. The area is composed of
350 points or spectra.
To correlate an optical image with the Raman profile, a at half maximum (FWHM) value method described in
small cross sectional piece of the nicotine patch was the Thermo Scientific Application Note 517181. Figure 2b
mounted vertically on the glass slide. shows the optical image of the physically cross-sectioned
sample. The optical image provides a visual representation
The Thermo Scientific™ OMNIC™ for Dispersive Raman of the layer thickness based on contrast and although it is
Software Suite, which includes Thermo Scientific™ comparable to the chemical image, some of the layers are
OMNIC™ Atlμs™ Software, was used for data collection, not easily identifiable.
processing, and analysis, including identification of the
layer composition and estimation of the layer thickness. Figure 3 shows the representative Raman spectra for
The multi-component spectral analysis was performed each layer in the Raman line depth profile. The chemical
using Thermo Scientific OMNIC Specta Software. composition of each layer was initially identified
by performing a library search against the Thermo
Results and discussion Scientific High-Resolution Raman Polymer Library using
A Raman line depth profile map obtained using the OMNIC OMNIC software.
software suite is shown in Figure 2a. In the figure, the depth
profile is displayed in a 2-D contour map and reveals the Layer 1, the backing is poly(ethylene terephthalate) or
changes in the Raman peaks (vertical lines) as the focal PET. Layers 2 and 4 are composed of microporous
point probes through the layers. The rainbow color scheme polyethylene (PE) that acts as a controlled release agent
of the contour map shows the intensity of the Raman peaks for nicotine, which is located at the ethylene/vinyl acetate
as color, with red indicating the highest intensity and blue copolymer (EVA) reservoir layer (layer 3). Layer 5 is a
the lowest intensity. A total of six layers were identified polyisobutylene (PIB) adhesive and PET. Layer 6, the
with thicknesses ranging from ~15 μm to ~75 μm. The release liner is also composed of PET.
thickness of each layer was estimated by the full width
Figure 5: X-Z correlation profile images generated by the reference Raman spectrum of each layer.
Conclusion References
Existing techniques for analyzing a multilayer polymer like 1. Guillory P., Deschaines T., Henson P. Confocal Raman
a transdermal nicotine patch are labor intensive, time- microscopy analysis of multilayer polymer films,
consuming, and would require destruction of the sample. Thermo Scientific Application Note 51718, 2008.
Using confocal Raman analysis, the analysis was quick,
and the sample integrity was preserved. The chemical
composition of each layer was determined using OMNIC
for Dispersive Raman and OMNIC Specta software suites.
OMNIC Specta Multi-Component Library Search identified
the constituents in layer 3, a mixture of nicotine and an EVA
co-polymer, which is easier than tedious manipulations
required by traditional search and subtract processing.
A B
Data processing through statistical analysis. It uses spectral covariance and
Peak area for pathlength correction option in the TQ factorial analysis to extract significant and relevant chemical
Analyst software was used to normalize Raman spectral information from sample spectra as factors, then correlate
intensities. The peak area of the CH2 bending mode at them with sample properties such as concentration,
1440 cm-1 was used for the normalization (1422-1452 cm-1 crystallinity and density. A total of 20 samples, a mix of
range, Figure 2A). An averaged two-point baseline pellets and films, were used as the calibration standards.
correction was used to account for baseline shifts/noise. Five additional pellet and film samples with density
values spread across the density range of the samples
Developing a PLS model for PE density determination were selected as the validation standards (Table 1). A
Partial Least Squares (PLS) algorithm11 from the TQ Analyst spectral range of 1400-1500 cm-1 with averaged two-point
software was used to develop a model for PE density correction for baseline (Figure 2B) was used in the method.
determination. PLS is a quantitative regression algorithm
Table 1: PE samples and their densities. *Rows highlighted in green are the samples used for validation.
Figure 3 shows the calibration results for PE densities important to note that the percent differences do not
obtained with the 3-factor PLS model. The inset is the exhibit any bias between pellets and films, indicating that
Predicted Residual Error Sum of Squares (PRESS) plot. the sample form (pellets vs. films) has no bearing on the
In the current case, a 3-factor model suffices as the model performance.
contribution from the 4th and 5th factors are negligible.
The calibration curve has a correlation coefficient of Density Determination of PE films
0.9914. The RMSE (Root Mean Square Error) values are Figure 4 demonstrates an example of applying the
0.00360 for the calibration samples and 0.00432 for the Raman spectrum and the PLS model to predict the
validation samples, respectively. The results are also density of a PE film sample. The predicted density is
summarized in Table 1. For all 25 samples, the calculated 0.9014 g/cm3, showing a good agreement with the actual
densities are within ±0.81% of the actual values. It is density of 0.9008 g/cm3.
Figure 3: Calibration results for PEs of different densities using a PLS quantitative analysis. The denotes calibration standards
and the + denotes validation standards. Inset is the PRESS plot for the PLS calibration. Three factors were used in the current PLS
calibration model.
Figure 4: Prediction of the density of a PE film sample using its Raman spectrum and the PLS model. The known density of the film is
0.9008 g/cm3 and the predicted density is 0.9014 g/cm3.
Figure 5A shows the Raman confocal depth profile of Conclusions
a clear transdermal nicotine patch. A total of 6 polymer Raman microscopy is a powerful analytical tool for PE
layers were identified, including two PE layers – PE density determination. Since PE chains in crystalline
layer 1 (part of the backing layer) and PE layer 2 (closer and amorphous domains exhibit unique Raman features
to the release liner). There are perceptible differences in in the CH2 bending region, a PLS model based on
the Raman spectra between the two PE layers (Figure 5B the Raman features in the 1400-1500 cm-1 region was
and 5C). Applying the PLS model, the densities are successfully developed. The model is applicable for both
determined to be 0.9150 g/cm3 for PE Layer 1 and pellet and film samples, showing a good agreement
0.9583 g/cm3 for PE Layer 2, placing PE layer 1 in the between actual and predicted density values. Applying
LDPE /LLDPE class and the PE Layer 2 in the HDPE the model to a real-world multilayer film containing two
class. The classification of the PE layers based on the PE layers, the predicted density values correctly place
predicted densities conforms to other reports: LDPE/ the two layers into their respective PE classes. More
LLDPE is used in the occlusive backing layer for its importantly, the confocal capability of Raman microscopy
flexibility whereas HDPE is used as the rate-controlling allows for in situ density determination of PE layers within
membrane as an integral part of the reservoir diffusion multilayer polymer films, without the need for tedious and
control machanism.12-13 While the exact densities of the challenging sample preparations required by many other
two PE layers are not available, the results presented techniques. The presented methodology should be of
here nonetheless demonstrates the advantage of using interest for PE manufacturers as well those who perform
Raman microscopy combined with the PLS method for failure analysis, reverse engineering, and polymer
density determination. The confocal capability of Raman composites development.
microscopy allows for in situ PE density determination in
multi-layer films without the need to isolate the individual Acknowledgement
PE layers. The authors would like to thank Wanda Weatherford at
Chevron Phillips Chemical Company for providing PE
pellet samples and Rajesh Paradkar at Dow Chemical
Company for providing PE film samples.
A B
Figure 5: (A) Confocal Raman line depth map of a multilayer polymer patch, showing the presence of two types of PE layers. (B) Raman
spectra of the two PE layers in full spectral range. (C) Raman spectra of the two PE layers in the CH2 bending and CH2 twisting region.
The insets in (C) show the calculated densities of the two PE layers using the 3-factor PLS model. PET = poly(ethylene terephthalate),
EVA = ethylene/vinyl acetate copolymer, PIB = polyisobutylene.
References
1. Piringer O.G. and Baner A.L., ed. Plastic Packaging: Interactions with Food and Pharmaceuticals. 2nd ed.
Weinheim: Wiley-VCH; 2008.
2. Polyethylene, The Essential Chemical Industry – online, http://www.essentialchemicalindustry.org/polymers/
polyethene.html, retrieved on 11/27/2017.
3. (a) ISO 1183-1:2012, Plastics – Methods for determining the density of non-cellular plastics − Part 1: Immersion
method, liquid pyknometer method and titration method; ANSI New York, NY. (b) ASTM D792-13, Standard test
methods for density and specific gravity (relative density) of plastics by displacement; ASTM International, West
Conshohocken, PA, 2013.
4. (a) ISO 1183-2:2004, Plastics – Methods for determining the density of non-cellular plastics – Part 2: Density
gradient column method. ANSI New York, NY. (b)ASTM D1505-10, Standard test method for density of plastics by
the density-gradient technique; ASTM International, West Conshohocken, PA, 2010.
5. ASTM D4883-08, Standard test method for density of polyethylene by the ultrasound technique; ASTM
International, West Conshohocken, PA, 2008.
6. Mieth A., Hoekstra E., and Simoneau C. Guidance for the identification of polymers in multilayer films used in food
contact materials: User guide of selected practices to determine the nature of layers, EUR 27816 EN, 2016; doi:
10.2788/10593.
7. Sato H., Shimoyama M., Kamiya T., Amari T., Sasic S., Ninomiya T., Siesler H.W., and Ozaki Y. Raman spectra of
high-density, low-density, and linear low-density polyethylene pellets and prediction of their physical properties by
multivariate data analysis, J. Appl. Polym. Sci., 2002, 86, 443–448.
8. Strobl G.R. and Hagedorn W. Raman spectroscopic method for determining the crystallinity of polyethylene, J.
Polym. Sci. B Polym. Phys., 1978, 16, 1181-1193.
9. Williams K.P.J. and Everall, N. J. Use of micro Raman spectroscopy for the quantitative determination of
polyethylene density using partial least-squares calibration, J. Raman Spectrosc. 1995, 26, 427-433.
10. Ibrahim M. and He H. Classification of polyethylene by Raman spectroscopy, Thermo Scientific Application Note
AN52301, 2017.
11. Thermo Scientific Product Overview. TQ Analyst Software Chemometric Algorithms, 2009.
12. Kearney C. J. and Mooney D.J. Macroscale delivery systems for molecular and cellular payloads, Nat. Mater.,
2013, 12, 1004-1017
13. Allen L. V. and Ansel H. C., ed. Ansel’s Pharmaceutical Dosage Forms and Drug Delivery Systems, 10th ed.
Lippincott Williams & Wilkins; 2013.
No. AN52959
Simultaneous rheology and Raman
spectroscopy during the melting and
recrystallization of polypropylene
Authors: Nathan C. Crawford and Figure 1:
David Drapcho The Thermo Scientific HAAKE
MARS XR RheoRaman system.
Introduction
Rheology is the study of the flow and deformation of
matter, including fluids to solid-like materials (and anything
in between). Rheological measurements are commonly
used to examine or induce bulk physical material
changes, such as melting, crystallization, gelation, and/
or polymerization etc. Raman spectroscopy, on the other
hand, is a vibrational spectroscopic technique that can
provide insights to these processes at the molecular
level, both chemically and morphologically. However,
until very recently1, Raman studies are often conducted In this note, a MARSXR RheoRaman system was used
ex-situ, whereby Raman spectra are acquired prior to to investigate the temperature-dependent melting and
and after the observed physical transformation. The true crystallization of polypropylene, as well as the isothermal
chemical/morphological changes driving these processes crystallization process. The melt and crystalline phase
are largely left uncaptured, leaving much room for data transitions of polymeric materials are commonly correlated
interpretation and speculation surrounding these dynamic with variations in viscous and elastic behavior during
physicochemical relationships. rheological analysis. In addition, these phase transitions
are often associated with spectral pattern changes
The Thermo Scientific™ HAAKE™ MARSXR (Figure 1) in characteristic Raman peaks during spectroscopic
is a fully integrated RheoRaman system that enables investigation. Measurements performed ex-situ are
simultaneous rheology and Raman spectroscopy often challenging to compare due to discrepancies
measurements. The seamless hyphenation of these two in temperature control, slight deviations in sample
techniques allows for real-time, in-situ measurements of composition, and differences in processing history. The
both physical and chemical/morphological properties with in-situ RheoRaman system, on the other hand, completely
great ease. This multimodal analytical tool offers several eliminates these discrepancies, allowing for a more valid
advantages compared to the traditional ex-situ approach. analysis of the melt and crystalline phase transitions from
First, data collection efficiency is greatly improved by both the macroscopic and molecular levels.
combining multiple one dimensional experiments into a single
multifaceted experiment. Secondly, because both techniques
are employed simultaneously, sample fidelity is preserved
and the transformation of the material is captured in real-
time. And finally, sample consumption is reduced, which
can be beneficial for new material/formulation development
where sample quantities are limited and/or expensive.
Materials and methods The iXR Raman spectrometer is free-space coupled to
Materials the rheometer with an optical train which uses two plane
Polypropylene (Ineos Olefins and Polymers, USA, R12C-00 mirrors to direct the incident laser into the RheoScope
random copolymer) pellets were used for this study. The module. Within the RheoScope, a series of mirrors directs
pellets were melted at 190 ºC on the rheometer to form a the laser beam into a 20x objective, where the laser light
continuous, disk-shaped specimen for testing. is focused through a 2 mm thick fused silica window
into the sample (perpendicularly to the flow or vorticity
Rheometer plane). Raman scattered light is collected in a 180º
Rheological measurements were performed using the backscatter geometry using the 20x objective, and back
Thermo Scientific™ HAAKE™ MARS™ III rheometer, into the spectrometer through the same optical train as
equipped with a 20 mm diameter stainless steel parallel- the incident laser (eventually to the spectrograph inside
plate rotor. In order to extract viscoelastic behavior of the spectrometer). Free-space coupling of the laser to the
the polypropylene, all measurements were conducted rheometer, and the iXR spectrometer design, allow easy
in the oscillatory mode. Oscillatory measurements were Raman excitation laser wavelength interchange to permit
performed at 1 Hz with a constant strain of 0.1%, while optimization of the laser wavelength to the sample (785 and
data were collected every 5 s. To soften the polypropylene 455 nm laser sets are also available).
and help it conform to the measuring geometry, all
samples were initially loaded at 190 ºC. For the melt-to- The sample is positioned between the silica window and the
crystallization phase transition study, the temperature was rotor geometry attached to the rheometer measuring head
decreased from 190 ºC to 30 ºC, at a rate of 5 ºC/min. For (Figure 2). The objective can be adjusted for interrogation
the isothermal crystallization studies at 138 ºC and 150 at different penetration depths within the sample, as well
ºC, temperature was rapidly decreased from 190 ºC to 10 as positioned at various radial locations along the optical
ºC above the target temperature. The temperature was slit (from the true center to outer edge of the sample). An
then slowly decreased until it reached the test temperature electrical heating element is positioned below the fused
(either 138 ºC or 150 ºC). The temperature was then silica window to provide temperature control during testing.
held constant for a maximum of 1 h (3600 s), while the Cooling was provided from a temperature-controlled
isothermal recrystallization process was observed. circulator with a 50:50 mixture of ethylene glycol and
water. All instrumentation is controlled through the Thermo
Spectrometer Scientific™ OMNIC™ and RheoWin software packages.
Raman spectroscopy measurements were
performed using the Thermo Scientific™
iXR™ Raman Spectrometer. The iXR
system employed a 532 nm, 10 mW
laser, a triplet spectrograph providing
Raman spectra over the range 3500 to
50 cm-1 Raman shift (Stokes) at 5 cm-1
resolution, and a CCD camera cooled
to -50 ºC. Alignment of the laser, Raman
scatter, and aperture selection within
the spectrometer were all software
controlled. The minimum exposure
collection time is 0.1 s. For the data
presented here, the exposure collection
time was 4 s and 2 sample exposures
were averaged per spectra collection.
Figure 2: Schematic diagram of the MARS XR RheoRaman system (showing side and
RheoRaman coupling
top views of the rheometer sample stage). The iXR Raman spectrometer is free-space
The Raman spectrometer (Thermo Scientific iXR Raman coupled to the MARS rheometer using plane mirrors that direct light into a 20x long-
Spectrometer) and rheometer (Thermo Scientific HAAKE working-distance objective. The objective focuses the incoming laser (green dashed
line) and collects the back-scattered Raman light (yellow) coming into and out of the
MARS III) are coupled together using the Thermo Scientific™
sample (which sits atop the rheometer stage). The optical path from the spectrometer
HAAKE™ RheoScope module (Figure 2). to the rheometer is enclosed in lens tubes (black). The optical slot in the bottom
heating plate permits passage of the laser beam and scattered Raman light.
The electric heater allows for a maximum temperature melt-to-recrystallization process2, while the spectral features
of 300 ºC and a minimum temperature of -5 ºC, with a in the 3050-2800 cm-1 range were integrated into a total peak
maximum heating/cooling rate of 10 ºC/min. An active area. These indicators (normalized helical chain vibration
electrical hood was also used to provide temperature peak height and overall CH stretching peak area) were then
control from above, eliminating the potential for a tracked throughout the melt- to crystalline-phase transition
temperature gradient within the sample (Figure 2). and overlaid with the in-situ rheology data (Figure 4).
The melt and recrystallization process of polypropylene
Results and discussion was probed rheologically using small amplitude oscillatory
The physicochemical and morphological relationships shear measurements (Figure 4), where the storage modulus
during polymer crystallization are of critical importance (G’) and loss modulus (G”) were measured as a function of
for polymer processing. Here we use the new MARSXR temperature. G’ and G” are measures of a material’s elastic
RheoRaman system to investigate the crystallization process and viscous behavior, respectively. A liquid-like material will
of polypropylene (PP). Representative Raman spectra for be more viscous than elastic (i.e., viscously dominated),
PP at three different temperatures are shown in Figure 3. In and as a result, G” will be greater than G’. Conversely, a
general, the spectrum at 53 ºC shows sharp peaks across solid-like material will display more elastic than viscous
the examined spectral range. These sharp spectral features behavior (i.e., elastically dominated), where G’ will be
suggest a high degree of conformational order, indicative of greater than G”.
semicrystalline and crystalline structures. As temperature
increased to 120 and 173 ºC, these peaks began to broaden
and merge together. The broadening of spectral peaks is
commonly ascribed to melt behavior in polymeric materials.
References
1. A.P. Kotula, M.W. Meyer, F. De Vito, J. Plog, A.R. Hight Walker and K.B. Migler, Review of Scientific Instruments, 87,
105105 (2016); DOI 10.1063/1.4903746.
2. R.A. Khafagy, Journal of Polymer Science: Part B: Polymer Physics, DOI 10.1002/polb
For Research Use Only. Not for use in diagnostic procedures. © 2017 Thermo Fisher Scientific Inc. All rights
reserved. All trademarks are the property of Thermo Fisher Scientific and its subsidiaries unless otherwise specified.
AN52959_E 06/17M
APPLICATION NOTE AN53002
In general, TAG molecules assume a tuning fork configuration and the TAG
“forks” assemble to form crystal lattice structures. During crystallization,
the TAG molecules slow down as the CB oil cools and the TAGs come to
rest in contact with one another, forming what are known as “sub crystalline
cells.”1 Once the sub-cells are formed, they are thermodynamically driven to
aggregate into larger and more stable crystalline structures.2 The self-assembly
of sub-cell structures and their further aggregation is governed by a balance
of intra- and inter-molecular interactions. Depending on the molecular level
packing and orientation of the TAGs, CB can form different types of crystal
lattice structures (or polymorphs), where some crystal structures are more
desirable than others. Overall, CB crystallization is a highly complex, multistage
process. Understanding the isothermal crystallization behavior of CB is vital for
improving chocolate manufacturing processes and maintaining product quality.
In this note, rheology coupled with in situ Raman Rheology
spectroscopy was used to examine the isothermal Rheological measurements were performed using a
crystallization of cocoa butter. Raman spectroscopy Thermo Scientific™ HAAKE™ MARS 60 Rheometer
is a highly sensitive, relatively fast, and nondestructive equipped with a serrated 35 mm diameter plate rotor
technique that can probe the molecular structure and at a gap height of 1 mm. The serrated plate was
conformation in both liquid and solid TAG assemblies, used to prevent slip at the sample-rotor interface. All
as well as intra- and inter-TAG interactions. With measurements were conducted in the oscillatory mode,
simultaneous Raman spectra and rheological data, with a frequency of 1 Hz and a constant strain of 0.1%.
molecular-level interactions and conformational shifts CB samples were loaded onto the rheometer at 60 °C
during the isothermal crystallization of CB were directly and allowed to equilibrate for 10 min to erase any crystal
correlated with the changes in bulk viscoelastic structures and/or shear history from sample loading.
properties, providing unique insight into the multifaceted After the equilibrium step, the temperature was rapidly
crystallization behavior of cocoa butter. decreased from 60 °C to 22 °C at a rate of 10 °C/min.
The temperature was then held constant at 22 °C for
Materials and methods 120 min, collecting data every 10 s.
Materials
Commercially available, organic cocoa butter
(Theobroma cacao) was acquired from Inesscents™
(Ashland, OR, USA).
A B
Figure 1. (a) The Thermo Scientific™ HAAKE™ MARSXR RheoRaman System. (b)
Schematic diagram of the MARSXR RheoRaman system (showing side and top views of
the rheometer sample stage). The iXR Raman spectrometer is free-space coupled to the
MARS rheometer using lens tubes and mirrors that direct light into a 20x objective. The
objective focuses the incoming laser (green dashed line) and collects the back-scattered
Raman light (yellow) coming out of the sample sitting atop the rheometer stage.
Although less intense than the C-H stretching region, Two well-defined features emerged at 1130 cm-1 and
approximately eight unique spectral features were 1063 cm-1 during the solidification process, which
identified in the fingerprint region (1000-1800 cm-1; originate from the symmetric and asymmetric C-C
Figure 4). When comparing the CB melt state to the stretching, respectively.4,5 In the melt phase, all C-C
crystalline phase, the most significant changes were stretching bands were relatively weak and broad due to
observed in the C–C stretching region (1000-1200 cm-1). the disordering effects of methyl gauche conformations.
Figure 4. The 1000–1800 cm-1 Raman spectral
range for melted and crystalline cocoa butter.
However, as the CB solidified, the backbone methyl Unlike the individual moduli, tan(δ) can be used to quantify
groups were ordered into the trans-conformation, overall brittleness of a material and is commonly used
signified by the emergence of the peak at 1130 cm-1. to assess glass transition behavior. In general, as tan(δ)
Therefore, in addition to the I2882/I2850 peak intensity ratio, becomes smaller, the more G’ deviates from G”, and the
the I1130/I2850 spectral marker was also used to track more brittle (or glass-like) the material becomes.
the crystalline-phase transition within CB via in situ
rheoRaman measurements. During the initial portion of the isothermal hold at 22 °C
from 0 to 5 min (immediately following the rapid decrease
Simultaneous rheology and Raman spectroscopy in temperature from 60 °C to 22 °C), both G’ and G”
(RheoRaman) increased as the CB transformed from a melted liquid to a
The melt-to-solid phase transition of cocoa butter was soft semi-solid (Figure 5a). This initial increase in modulus
probed rheologically using small amplitude oscillatory is most likely due to a delay between the set temperature
shear measurements (Figure 5a), where the storage and the internal temperature of the loaded sample. Once
modulus G’ and loss modulus G” were measured as a the sample had reached thermal equilibrium and was
function of time at the isothermal temperature of 22 °C. at the isothermal set point of 22 °C, the moduli were
G’ and G” are measures of a material’s elastic and relatively stable from 10 to 25 min. From 25 to 50 min,
viscous behavior, respectively. A liquid-like material will however, both G’ and G” begin to gradually increase and
be more viscous than elastic (i.e., viscously dominated), then from 50 to 80 min, the moduli rapidly increased,
and as a result, G” will be greater than G’. Conversely, a where G’ and G” increased by approximately 5 and
solid-like material will display more elastic than viscous 4 orders of magnitude, respectively. The exponential
behavior (i.e., elastically dominated), where G’ will be increase in the moduli indicates a solidification process,
greater than G”. The overall magnitudes of G’ and G”, where the CB transformed from a pliable semi-solid to
as well as their relative difference in magnitude, often a more robust, hardened solid. At 80 min and beyond,
reported as the ratio of G”/G’, determines the general growth in the elastic modulus slowed and eventually
viscoelasticity and overall resistance to deformation for a plateaued, showing no further significant change past
given material. 100 min. The viscous modulus, however, reached a
slight plateau from 80 to 100 min and then proceeded to
The ratio of G”/G’ (plotted on the right y-axis of Figure 5a) gradually decrease from 100 min and beyond.
is commonly used to track viscoelasticity of a material:
During the increase in G’ and G”, a rapid decrease in
the loss factor tan(δ) was observed from ~65 min and
beyond (Figure 5a, right y-axis). The decrease in the loss
where δ is the phase angle defined as the shift or lag factor indicates a deviation in overall magnitude between
between the input strain and resultant stress sine waves G’ and G”. As the CB hardened, the increase in G’
(or vice versa) during an oscillatory shear measurement. exceeded the increase in G”, triggering the decrease in
The term “tan(δ)” is often referred to as the loss or damping tan(δ). At the end of the 120 min isothermal study, G’ was
factor. Values of tan(δ) less than unity indicate elastically more than a full order of magnitude greater than G” and
dominant (solid-like) behavior, while values greater than the loss factor was approaching 0.01, indicating the CB
unity indicate viscously dominant (liquid-like) behavior. had transitioned into a brittle glass-like solid.
Raman spectral markers, on the other hand, are indicators
A of crystal formation. Thus, the time delay between the
rheology and Raman profiles suggests that CB first hardens
into an amorphous solid, followed by a transformation from
an amorphous to a crystalline solid. This morphological
transformation was signified by the subsequent increase
in the Raman band intensities associated with crystal CB
structures (the 1130 and 2882 cm-1 peaks). The temporal
separation of the rheological and Raman spectral profiles
indicates a clear distinction between bulk hardening of the
CB and the formation of crystalline domains.
Conclusions
Simultaneous rheology and Raman spectroscopy
measurements were used to examine the isothermal
crystallization of cocoa butter. This multimodal analytical
technique allowed the bulk mechanical properties of cocoa
Figure 5. (a) Rheology: G’ and G” (filled and open circles, respectively;
plotted on the left y-axis) and tan(δ) (plotted on the right y-axis) and butter (G’, G”, and tan(δ)) to be directly correlated with
(b) Raman: the I1130 /I2850 (left y-axis, green) and I2882/I2850 (right y-axis, conformational changes at the molecular level (νas(CH2)
black) peak intensity ratios for CB during isothermal crystallization
mode at 2882 cm-1 and the νs(C-C) mode at 1130 cm-1)
at 22 °C. The vertical dashed line at 45 min indicates the increase of
G’ and G”, while the dashed line at 65 min indicates the decrease in in real-time. After rapid cooling (10 °C /min) and at an
tan(δ) and increase in the Raman ratios. isothermal temperature of 22 °C, there was a noticeable
time lag between the rheological response (G’ and G”)
The observed rheological behavior was further confirmed and the Raman spectral profiles. The observed time delay
using simultaneous Raman spectroscopy (Figure 5b). indicates that CB crystallized by first hardening into an
Initially, both the I1130/I2850 and I2882/I2850 peak intensity amorphous solid, manifested by a sharp increase in G’ and
ratios remained unchanged during the first ~65 min of the G” while the Raman features remained unchanged. The
isothermal study. Then a sharp increase of the I1130/I2850 and amorphous solid then underwent a morphological transition
I2882/I2850 ratios began at ~65 min, indicating the formation to form a crystalline solid, signified by the increase in
of crystal structures within the CB. As the CB further Raman features associated with crystal CB structures
crystallized, both spectral markers continued to increase (1130 and 2882 cm-1). Without coupling these two separate
from 65 to 100 min. Beyond 100 min, the growth in both analytical techniques, the observed amorphous-solid
Raman features had subsided and the peak intensity ratios to crystalline-solid transformation would have been left
began to stabilize. undetected. Alone, each technique suggests a single-
stage process, however, only when the two techniques are
Overall, the rate of increase in the 1130 and 2882 cm-1 coupled is the multi-phase crystallization process revealed,
spectral ratios were similar to the rate of change for both G’ further exemplifying the unique analytical capability
and G” (i.e., they increased with similar slopes). However, unleashed by hyphenating rheology with in situ Raman
there was a noticeable 15-20 min lag between the observed spectroscopy. While this work focusses on the isothermal
increase in G’ and G” and the rise of the Raman intensity crystallization of CB, the underlying principles applied here
ratios. The sharp upturn in G’ and G” indicates an increased should be applicable for a wide range of material processes
resistance to deformation (i.e., a bulk hardening of the including gelation, polymerization, curing behavior, as well
CB), signaling the start of the solidification process. The as other shear-induced phenomena.
References
1. K. Sato, Crystallization of Lipids: Fundamentals and
Applications in Food, Cosmetics, and Pharmaceuti-
cals. Hoboken, NJ: John Wiley & Sons, 2018.
2. S. Bresson, D. Rousseau, S. Ghosh, M. El Marssi,
and V. Faivre, Raman spectroscopy of the polymor-
phic forms and liquid state of cocoa butter, Eur. J.
Lipid Sci. Technol. 113, 992–1004, 2011.
3. R. G. Snyder, H. L. Strauss, and C. A. Elliger, C–H
stretching modes and the structure of n-alkyl chains.
1. Long, disordered chains, J. Phys. Chem. 86,
5145–5150, 1982.
4. R. J. Meier, Studying the length of trans confor-
mational sequences in polyethylene using Raman
spectroscopy: A computational study, Polymer. 43,
517–522, 2002.
5. M. Zheng and M. Du, Phase behavior, conformations,
thermodynamic properties, and molecular motion of
multicomponent paraffin waxes: A Raman spectros-
copy study. Vib. Spectrosc. 40, 219–224, 2006.
Abstract
This article discusses basic practical aspects of analytical instrument performance and
experimental design that should be taken into consideration when characterizing microscale
objects using a Raman microscope. Proper instrument alignment, optical objective magnification,
confocal aperture, and sampling step settings must ensure sufficient spatial resolution and
measurement precision to discriminate a Raman signal of the object from a surrounding matrix.
The essential relationships between the spectral measurement parameters are considered
theoretically and illustrated experimentally.
Read online ›
Spectroscopy, 2016, Volume 31, Issue 11, p 40–45
5 School of Electrical and Electronic Engineering, Nanyang Technological University, 639798, Singapore.
Abstract
Raman microspectroscopy provides for high-resolution non-invasive molecular analysis
of biological samples and has a breakthrough potential for dissection of cellular molecular
composition at a single organelle level. However, the potential of Raman microspectroscopy
can be fully realized only when novel types of molecular probes distinguishable in the Raman
spectroscopy modality are developed for labeling of specific cellular domains to guide
spectrochemical spatial imaging. Here we report on the design of a next generation Raman
probe, based on BlackBerry Quencher 650 compound, which provides unprecedentedly high
signal intensity through the Resonance Raman (RR) enhancement mechanism. Remarkably,
RR enhancement occurs with low-toxic red light, which is close to maximum transparency in
the biological optical window. The utility of proposed RR probes was validated for targeting
lysosomes in live cultured cells, which enabled identification and subsequent monitoring of
dynamic changes in this organelle by Raman imaging.
Read online ›
Sci. Rep. 6, 28483; doi: 10.1038/srep28483 (2016)
Tracking microplastics in the environment via
FTIR microscopy
Authors
Michael Bradley, Suja Sukumaran, Steven Lowry, Stephan Woods
Thermo Fisher Scientific, Madison WI, USA
Abstract
Microplastics are particulates, roughly 20–1000 μm in size, originating from materials such as
clothing, abrasive action on plastics, or engineered microbeads as found in some exfoliating
cosmetics. The microplastics enter aquifers where the particles can be consumed by filter
feeders. Microplastics are chemically stable, giving them a long lifetime in the environment and
making excretion or digestion difficult. Analytically, the size and polymeric nature of microplastics
makes Fourier transform infrared (FTIR) microscopy an ideal tool for detection and identification.
Standard analyses typically start with a filtration step, extracting the material from the matrix. The
analysis can proceed directly on the dried filter without further sample preparation. This simplicity
in both sampling and analysis enables the rapid assessment of microplastic encroachment and
can assist in the development of remediation techniques. We show examples from both prepared
and field samples using microattenuated total reflection (ATR) FTIR.
Read online ›
Spectroscopy, 2017, Special Issues Volume 32, Issue 8, p 17–23